Vous êtes sur la page 1sur 52
CHAPTER 13 Centrifuge modelling of soil behaviour 13.1 Introduction Physical modelling of soil behaviour has always played a pivotal role in helping the designer acquire a better understanding of the actual behaviour under similar stress conditions in the field. In this respect, laboratory tests such as triaxial, consolidation and shear box tests are still used extensively and do provide reliable data provided that adequate testing procedures are adhered to. Centrifuge testing of soils constitutes another (recent) development in the field of physical modelling, which has taken off quite rapidly. More and more research laboratories are equipped with centrifuge centres, and the research outcome in this field is becoming widely available so that comparative studies can be undertaken, leading to the development of ever more sophisticated instrumentation equipment. As in the case of triaxial testing, for instance, a centrifuge test is undertaken on a small size sample of soil, referred to as the model, in a way that stress conditions corresponding to a particular event (such as subjecting a pile to a lateral load for example), or to a particular process (as in the case of the execution of an excavation, or the construction of an embankment) are recreated via an inertial acceleration field. Under such loading conditions, the behaviour of the mode/ should be, in theory, a replica of that of the actual soil (often referred to as the prototype), when subjected to a similar state of stresses. Notwithstanding the practical difficulties related to model preparation and instrumentation, centrifuge modelling is regarded as a valuable means of testing that enhances markedly the understanding of the physical behaviour of soils under complex static or dynamic stress fields (see for instance the paper by Schofield (1980)). Examples vary widely and include (classical) problems such as the stability of slopes, retaining structures, embankments, foundations and tunnels, as well as heat transfer (that is conduction and 692 CENTRIFUGE MODELLING OF SOIL BEHAVIOUR convection), diffusion (i.e. consolidation), seepage, earthquakes, wave loading, contaminant transport, freeze/thaw and the effects of deep mining. 13.2 Modelling principles, stress similarities The basic principle of centrifuge testing consists of creating stress conditions which are similar to those applied through gravity in the field to a prototype, in or on a model, with dimensions which are much smaller than those of the prototype. This can be achieved by placing the model in a basket at the end of a centrifuge boom, then subjecting it to an inertial acceleration field. The main features of a centrifuge equipment are depicted in figure 13.1, corresponding to an Acutronic 680 centrifuge, capable of developing a maximum acceleration of 200g (g being the acceleration due to gravity), and whose technical details can be found for instance in the book Centrifuge 88 edited by Corté (1988). basket in full swing ounterweight centrifuge boom aerodynamic enclosure drive system A ce Figure 13.1: Acutronic 680 centrifuge. In the following analysis, the subscript m refers to the model and the subscript p is used in conjunction with the prototype. MODELLING PRINCIPLES, STRESS SIMILARITIES 693 Thus, assuming that all soil properties, including those of the porewater, are identical for the prototype and for the model, and referring to figure 13.2, it is seen that, for a model with a dimension R/n (n being an integer), placed at a distance R (which can be the effective radius of a centrifuge, for instance) from the centre of rotation, and subjected to an acceleration Ng (ie. N times the acceleration due to gravity g), the centrifugal stress Om is, in theory, similar to the vertical stress oy, due to the self weight of soil in a prototype having a dimension NR/n. on angular velocity @ —_ force Figure 13.2:Effect of inertial forces on stress similarities between model and prototype. Accordingly, if the respective dimensions of the model and prototype were: hin=% and hyp =NE then the vertical stress Gym at a depth Am in the model, induced by an acceleration Ng will be: Sym = Npghm (13.1) Similarly, the vertical stress oy, at depth Ap in the prototype is: vp = pghp (13.2) 694 CENTRIFUGE MODELLING OF SOIL BEHAVIOUR p, in both equations, being the soil density. If the stresses are similar, then clearly the dimensions of the model and prototype are such that: hp him = = . m= (133) In practice however, the similarity between Gym and Oyp is affected to a certain degree by the negative centripetal (or inertial) forces applied to the model due to the variation of the angular velocity with the radius r. Referring once more to figure 13.2, it can be seen that the stress Oy», at the base of the model can be evaluated as follows: R 2 = andr = a? (an=1) 134) ove Tee porrar = pas n? (13.4) Moreover, the vertical stress due to gravity at the base of the prototype is: Ov = penE (13.5) Consequently, for these two stresses to be similar, equations 13.4 and 13.5 must be equal. Whence: N, Om =Oy > oesl-$ (13.6) It can be shown that if the stresses Oym at a depth R/n and Gyp at a depth NR/nare similar, then the maximum stress difference (i.e. error) defined as: Ow —5. AS= (13.7) occurs at a depth R/2n in the model and NR/2n in the prototype (see for instance Taylor (1995)). Thus, evaluating the stresses at these respective depths, it follows that: -R (1-1/2n) R ae f pordr = po? (4n—3) (13.8) R(1-In) 8n2 MODELLING PRINCIPLES, STRESS SIMILARITIES 695 and R Ow = pEN5- (13.9) Consequently, it is easy to show that the error corresponding to equation 13.7 is such that: 4nNs As=— 5 (13.10) @7*R(4n-3) Hence, substituting for the quantity Ng/o?R from equation 13.6 in equation 13.10 and rearranging, it follows that: 1 as=py (13.11) Moreover, assuming there is similarity of stresses at two-thirds model depth as depicted in figure 13.3, then the graph corresponding to the maximum error AS calculated from equation 13.11, and shown in figure 13.4, indicates that for an n value larger than 10, the ensuing error (i.e. the shaded area in figure 13.4) is smaller than 3% and therefore negligible. model prototype 2 2N Am Figure 13.3: Scaling errors due to inertial forces in the model. In practice, a value n = 10 is typical in geotechnical centrifuge testing, so that the model height corresponds to one-tenth of the effective centrifuge radius (measured from the centre of rotation to one-third the model depth). 696 CENTRIFUGE MODELLING OF SOIL BEHAVIOUR 5 10 IS 20 n Figure 13.4: Error magnitude related to model dimensions. 13.3 Scale effects The previous analysis of stress similarities between model and prototype serves as a reminder of the fact that physical (especially centrifuge) modelling of soils is affected by the reduction in dimensions of the tested volume of soil. The scale effects are not, however, limited to the model size, since two more phenomena related to centrifuge testing can, in principle, cause the behaviour of the model to be appreciably different from that of the prototype of an identical soil. These two phenomena are associated with (a) the particle size and (5) the rotational acceleration field. It has been established previously (refer to equation 13.3) that the stress similarity necessitates the use of a model with dimensions N times smaller than the prototype, with Ng being the acceleration to which the model is subjected in the centrifuge. Moreover, an equivalent dimensional analysis indicates that, for the same soil, the ratio d/L (d being the average grain size of the soil, and Z a typical boundary dimension) must, in theory, be identical for both the prototype and the model, so that: dn _ dp Lm Ep Substituting for the length Ln, is seen that: d, am = hm from equation 13.3, and rearranging, it (13.12) SCALE EFFECTS 697 Equation 13.12 therefore indicates the need to use a model comprising a soil with an average grain size N-times smaller than that of the same soil used for the prototype. Manifestly, this conclusion has, in practice, the potential of generating problems since, for instance, a model tested at an acceleration of 100g, constituted of fine sand that has in the field an average grain size of dy =0.1mm, must have an average grain size of dy/100 = 0.001 mm which, according to table 1.1 of section 1.2 corresponds to a clay, with all the implications related to the void ratio, permeability and stress-strain behaviour. Although no simple answer can be provided as to how the grain size might affect in any appreciable way the behaviour observed during a centrifuge test (see for example Ovesen (1979), Bolton and Lau (1988), Tatsuoka et al. (1991)), it seems logical, though, that the grain size effect decays with decreasing grain dimensions, so that while it appears necessary to reduce the grain size in proportion to the centrifuge acceleration when testing a gravel, the adverse effects of not doing so when testing a fine sand, a silt or a clay appear to be minimal. Evidently, these effects need be (or are already in the process of being) investigated thoroughly, and thus an engineering judgement has to be made in conjunction with the soil used for a model and the type of problem to be physically simulated. Therefore some caution must be exercised, so that the above analysis and guidelines are not used in a mechanical way. Let us now establish the effects of the rotational acceleration field, better known as the Coriolis effects. In this respect, it is easier to establish all acceleration components related to a soil element, rotated at a steady velocity V, much in the way that a model contained in a centrifuge basket is subjected to an acceleration Ng. Thus, with reference to figure 13.5, it is seen that the radial co-ordinates of point A, a distance r far from the centre of rotation, are as follows: X=rcos0 (13.13) Y=r sin@ The components of velocity are therefore: 698 CENTRIFUGE MODELLING OF SOIL BEHAVIOUR aK tt nsing® Vy = 4 = cos 00 —rsin8 7, = < ging Eo Vy = 9, = sin0F +rcos07, whence the following accelerations: de dt dt #Y _ ing ft ad ea ae in Ora ip + 200800 a mt +rcosOTy 2 7 ad ex = cos 0 Pr 2 sin 9x2 —rsino#2 — rcos0(®) -rsin (2) , Figure 13.5: Radial co-ordinate system. (13.14) (13.15) Moreover, when the local co-ordinates system (x, y) is used, then with reference to figure 13.6, the position of A is such that: x=C y=Cr-r where C and C2 are constants. (13.16) SCALE EFFECTS 699 Figure 13.6: Local co-ordinate system. Referring to figure 13.6, it is seen that: hy =r cos@sin® =X sind hy =h, = Ycos0 &; =X cos0 &) =Ysin0 and r=€i+&2 Accordingly, equations 13.16 can now be expressed as follows: x=Ci-hyt+hy y=C2-E1-&2 and so, inserting for the quantities h), hy, &; and &, it follows that: x=C;-X sin@+Ycos6 (13.17) y=C2-X cos0—Y sind 700 CENTRIFUGE MODELLING OF SOIL BEHAVIOUR The accelerations, expressed in the local co-ordinates system can then be derived: @&: &X aoe (2) sino +(“) cose e ae (2) cos0- (& ¥) sine Thus, substituting for the quantities derived in equation 13.15 and rearranging: a dt dt dt fy _ dr (#)’ ae ae "Na Equations 13.18 yield the four components of acceleration: 74 , 20 (13.18) : A &0 - the horizontal shaking: T,= ie 2. - the vertical shaking: v= ar dt 2 2 - the inertial acceleration: Tj=r (®) =ors £ where @ is the angular velocity Te =e - the Coriolis acceleration: dt dt It is noticeable that the first two components [, and Ty only apply to dynamic models (to simulate quakes for instance). Also, the Coriolis acceleration I’. which translates the velocity v = dr/dt of a particle within the model, relative to the velocity of the model centrifuge in flight = rd0/dt, can be rewritten as follows: Te =2v@ =2v¥ (13.19) SCALING LAWS 701 In practice, it is accepted that, for slow moving particles, in other words for a steady model flight where no shaking takes place, the error due to Coriolis acceleration is negligible for ratios ',/T'; < 0.1, meaning: v<0.05V (13.20) However, for faster particles, the path of a moving particle within a model becomes curved, with a radius of curvature r- defined as follows: 2 rear (13.21) Substituting for ', from equation 13.19, it is seen that: xis v = ap (13.22) Since r represents the effective radius of centrifuge (refer to figure 13.5), equation 13.22 implies that the curvature effect becomes less significant when r, >r. Accordingly, it is suggested that for fast particles (such as in a blast simulation for instance), the errors related to Coriolis acceleration are no longer appreciable as long as the velocities in equation 13.22 are such that: v>2V (13.23) Thus, both equations 13.20 and 13.23 yield the range within which Coriolis effects must be taken into account: 0.05VT-') in a model is N? times smaller than that occurring in the prototype. Diffusion problems The governing parabolic equation of a consolidation process (ie. porewater pressure dissipation with time) can be written as: 2. otaaet (13.37 The solution to the above equation contains the following time factor written in the case of the centrifuge model as follows: 2 m 1 Ty = 0,4 (13.38) di; SCALING LAWS 705 where L represents the drainage path and 1 is the time, the coefficient of consolidation c, being identical for both prototype and model. Substituting for tem and Lm from equations 13.32 and 13.25 respectively, then cancelling appropriate terms: 1 ton = stp (13.39) This equation, being identical to equation 13.32 in the case of seepage, indicates that the same stage of consolidation occurs N? times faster in a centrifuge model than in a prototype. Reynolds number The Reynolds number is of particular interest in centrifuge modelling, because its value is linked to the flow regime. It is well established that for soils, the regime of flow is assumed /aminar as long as the Reynolds number is kept smaller than 10 (refer to section 3.1). Let us examine the effect that this condition has on the centrifuge model. The dimensionless Reynolds number corresponding to the centrifuge model is defined as: Rem = md (13.40) where vm represents the model seepage velocity, d is the average diameter of soil particles and 1 is the kinematic viscosity of water. Since d and pare identical for both prototype and centrifuge model, equation 13.40 can then be rearranged after introducing the scaling law for seepage velocities from equation 13.30: Rem = vp = NRep (13.41) which implies that the Reynolds number is N times higher in a centrifuge model than in a prototype. Consequently, in order to maintain a /aminar regime of flow in the model, provisions must be made to keep the Reynolds number smaller than 10. It is of interest to note that, according to equation 13.41, an identical Reynolds number for both model and prototype can be achieved by using 706 CENTRIFUGE MODELLING OF SOIL BEHAVIOUR for the centrifuge model: (a) a soil characterised by an average grain diameter that is NV times smaller than the average diameter in the prototype or (6) a fluid N times more viscous than water. Clearly the first suggestion is unworkable, however, a variety of oils with differing viscosities can be used as a pore fluid in the model instead of water. One has to bear in mind, though, that the use of a pore fluid other than water may cause the surface properties of the solid particles to change, which may, in turn, affect the behaviour of the model. 13. Dynamic models Dynamic events such as quakes can be simulated in a centrifuge, and the scaling laws related to the corresponding models are derived in a manner similar to that used in conjunction with static models. Figure 13.7 depicts a typical horizontal shear wave generated by an earthquake, which is represented by the following differential equation: x= asinot (13.42) where x represents the cyclic motion, a is the amplitude of the motion, and @ corresponds to the angular velocity. Obviously, both velocity and acceleration can be derived from equation 13.42: & =awcos@t (13.43) £x — _aesinwt (13.44) a As the quantity am? in equation 13.44 defines the magnitude of acceleration, and because the centrifuge model is subjected to an acceleration N-times larger than that of the prototype, it is clear that: 2 2 AmOm = Nap®p (13.45) also, according to equation 13.25: am = 5 (13.46) SCALING LAWS 707 Thus inserting am into equation 13.45, and rearranging, it is easy to establish that: @m = Nop (13.47) and because the angular velocity is related to the motion frequency, it follows that: o=2nf > fn=Nfp (13.48) More importantly perhaps, the quantity aw in equation 13.43, which corresponds to the magnitude of velocity, is identical for both model and prototype. In fact, substituting for am and @m from equations 13.46 and 13.47 respectively into equation 13.43, it is seen that: Am®m COS(Wmt) = Ap @p COS (Np f) (13.49) Because the velocity magnitude corresponds to the ratio of a length to a time, and knowing that in a centrifuge model the length is reduced by a factor N (see equation 13.25), equation 13.49 therefore implies that the time for dynamic models is reduced by a factor N as opposed to a factor N* for seepage or diffusion problems (refer to equations 13.31 and 13.39). t os a Figure 13.7: Vertically propagating shear wave. Example 13.1 A model with a volume V,=3 x 107m} is subjected to 10 cycles at a frequency fm =100Hz and with an amliptude a» = 1.5mm, while being 708 CENTRIFUGE MODELLING OF SOIL BEHAVIOUR accelerated in a centrifuge at 100g. Obviously, the duration of shaking is calculated in a straightforward way: tm = 10/100 = 0.1s. The magnitude of acceleration due to the shaking can be estimated as follows: mm = Am(2fm)” = 1.5 x 103 x 4? x 104 = 592 mis? = 60.3¢ This test simulates an earthquake in a prototype, having a volume: Vp =N? Vm = 30, 000m? subjected to 10 cycles at a frequency: fp = ta =1Hz, and with an amplitude: ay = Nam = 0.15 m. § g The duration of the earthquake is such that: ¢) = Ntm = 10s, and the magnitude of acceleration in the prototype is, with reference to figure 13.8: Figure 13.8: Earthquake accelerations. ay(2nfy)” = 0.15 x 4n? x 1 = 5.92 mis? = 0.6g Assuming that, during the shaking, the steady centrifuge velocity is V = 30 mis, then the error due to Coriolis acceleration at maximum velocity of shaking, corresponding to the ratio of the inertial acceleration I, = V2/r to the Coriolis acceleration, =2vV/r is estimated as follows: V=Am2M fm = 1.5 x 20 x 107! = 0.94 mis whence: 2x0.94 30 et Sle =2v es =F 0.062 SCALING LAWS 709 Obviously, in this case, the error due to Coriolis acceleration is only 6.2%, and is therefore negligible. The scaling laws for different physical quantities are summarised in the following table in the case of a model subject to an acceleration N.g. quantity scaling law scaling factor acceleration gn =Ngp Nn! mass, density Pm =Pp 1 stress om =Op 1 strain Em =Ep 1 velocity Vin =Vp 1 temperature Om =O, 1 length Lm =L,/N N time (static event) tm = tp/N? N? time (dynamic event) tm = tp/N N displacement, amplitude dm =dyI/N N unit weight Ym =p Nt Srequency fm =Nfp No hydraulic gradient im =Nip Nt seepage velocity Vm =Nvp No Reynolds number Rem =NRep N! heat flux hum =Nhy Nt seepage flow per unit length |qy = qp/N? N? total seepage flow Qn — =Qp/NP Ni diffusion (consolidation) ton = tep/N? N? heat transfer(conduction, (60/2t),, = (8/61) ,/N? N convection) 710 CENTRIFUGE MODELLING OF SOIL BEHAVIOUR 13.5 Practical aspects of centrifuge modelling Every engineer or researcher would recognise that the outcome of a centrifuge test in geotechnics depends on the meticulous model preparation prior to testing, which is, alas, time consuming. In this respect, the paper by Phillips (1995) constitutes excellent reading as it describes in some detail the potential pitfalls related to model preparation, testing and monitoring. The one aspect that must always be considered first in relation to centrifuge testing is safety, Centrifuges such as the one depicted in figure 13.1 are powerful machines that must be handled by adequately trained staff. Furthermore, centrifuge testing is unquestionably a multi-disciplinary activity, since the modeller has to be at least conversant in (if not knowledgeable about!) mechanical engineering, electronics equipment, control systems and data acquisition, before even contemplating interpreting the test results from a geotechnical viewpoint. No wonder that such tests are generally intensively resourced, therefore expensive and time consuming. Manifestly, undertaking a centrifuge test implies spending the majority of time and effort on the careful preparation of the model. This includes the following. - The selection of the appropriate container depending on the type of soil tested and the type of behaviour to be simulated (static or dynamic) with specific boundary conditions requirements that need to be fulfilled. For instance, modelling a static event such as the consolidation of a clay layer requires the walls of the container to be ideally frictionless, whereas the simulation of the creep behaviour of a frozen soil necessitates a strict thermal control within the container. ~ The selection of the appropriate soil conditions for the model, in particular, the restrictions related to the flow of pore fluid and the size of the model must always be taken into account. It was established earlier that the Reynolds number in the model must be kept below 10 for laminar flow conditions to prevail. PRACTICAL ASPECTS OF CENTRIFUGE MODELLING 711 This requirement is most likely to be fulfilled through the use of a pore fluid that is N times more viscous than water in conjunction with the model (Ng being the acceleration to which the model is subjected). Also, in order to minimise the scaling errors due to inertial forces (refer to figures 13.3 and 13.4), the model height must be kept to within one-tenth of the effective centrifuge radius (measured from the centre of rotation to one-third the model depth). On the other hand, the model may have to be protected against any temperature changes or any air movements that can potentially be generated within the aerodynamic enclosure (see figure 13.1) during testing. - The reconstitution of the model under laboratory conditions must allow for the effective stress profile to be recreated in the case of cohesive soils, or for the change in soil behaviour to be taken into account when interpreting test results. In particular, remoulded samples of clay and silt can be created from a slurry by tamping. Alternatively, the slurry can be consolidated in the centrifuge, in which case, care must be taken not to induce a differential consolidation of the slurry mass through the generation of preferential drainage paths due to the high pore pressure within the soil mass (refer to section 5.6). For granular soils, sophisticated techniques such as tamping and pluviation can be used to create a density controlled model. - Instrumentation and actuation of the model can potentially create problems since neither the instruments, nor the actuator are scaled. Pore pressure, total stress and displacement transducers, as well as thermocouples, that are buried within the model should be placed in a way that minimises any reinforcement effect of the soil. Also, there is a need to insulate strain gauges and lead wires embedded in the model. The restrictive effects on the model behaviour of the actuator (i.e. the system that sets off the model to simulate the behaviour of the prototype) must be reduced to a minimum. - A data acquisition system that can store a large amount of data in a short period of time is needed to conduct a centrifuge test successfully. In this respect, sophisticated systems combining the latest electronics and digital technologies are commercially 712 CENTRIFUGE MODELLING OF SOIL BEHAVIOUR available. These points, important though they may be, are only an apercu of the acumen, hard work, and vision required by a modeller to conduct a sophisticated and technically challenging test such as a geotechnical centrifuge test. A thorough analysis which tackles more detailed practical aspects of centrifuge modelling can be found in Phillips (1995). References Bolton, M. D. and Lau, C. K. (1988) Scale Effects Arising from Particle Size. Centrifuge 88 (ed. J. F. Corté). A.A. Balkema, Rotterdam, pp. 127-134. Corté, J. F. (Ed.) (1988) Centrifuge 88. A.A. Balkema, Rotterdam. Ovesen, N. K. (1979) The scaling law relationship. Panel discussion. Proceeding of the 7th European conference on Soil Mechanics and Foundation Engineering, Brighton, 4, pp. 319-323. Phillips, R. (1995) Centrifuge modelling: practical considerations. Geotechnical Centrifuge Technology, Blackie, London, pp. 34-60. Schofield, A. N. (1980) Cambridge geotechnical centrifuge operations. Géotechnique, 20, pp. 227-268. Tatsuoka, F., Okahara, M., Tanaka, T., Tani, K., Morimoto, T. and Siddiquee, M.S. A. (1991) Progressive failure and particle size effect in bearing capacity of a footing in sand, ASCE Geotechnical Engineering Congress, Vol. 2 (Geotechnical special publication 27), pp. 788-802. Taylor, R. N. (1995) Centrifuges in modelling: principles and scale effects. Geotechnical Centrifuge Technology, Blackie, London, pp. 19-33. CHAPTER 14 Finite element modelling in geotechnics 14.1 Finite element modelling This section aims at presenting the numerical modelling used in conjunction with geotechnical problems from a practical perspective. As such, it is not intended to develop in detail the mathematical formalisms of the finite element method, although its working will be explained succinctly. A thorough presentation of such a method of analysis is widely available, and reference should be made to Zienkiewicz and Taylor (1991), Bathe (1982, 1996), Hughes (1987), Stasa (1985), Griffiths and Smith (1991), Smith and Griffiths (1998), Reddy (1993) and Owen and Hinton (1980). The finite element method is one of the most powerful approximate solution method that can be applied to solve a wide range of problems represented by ordinary or partial differential equations. The power of such a method derives from the fact that it can easily accommodate changes in the material stiffness which is evaluated at element level as will be explained shortly. Also, it allows for different boundary conditions to be applied in such a way that an acceptable global approximate solution to the physical problem can be achieved. Considering that closed form solutions cannot be elaborated for a large number of complex physical problems, due to the impossibility of satisfying the boundary conditions related to the corresponding equilibrium equations, the finite element method therefore provides an ideal alternative (approximate) solution method. In its simplest form, the finite element method consists of: - dividing a given structure or domain into a number of elements (hence the name finite elements); this process is known as discretisation. Elements are connected by nodes at the corners and sometimes at the sides as well; 714 FINITE ELEMENT MODELLING IN GEOTECHNICS - modelling the behaviour of the unknown variables at different nodes through the use of appropriate interpolation polynomials, better known as shape functions. The shape, size, number and type of elements depend on the type of structure or domain, and also on the precision required in the solution, these points being further elaborated below. In all cases, finite element modelling invariably leads to a matrix formulation that depends on the nature of the physical problem to be solved. Hence, steady-state problems (i.e. problems for which the unknown variables are independent of time) always consist of solving (numerically) the following matrix relationship: [K]{U} ={F} (14.1) where [K] represents the global stiffness matrix (that includes the material properties), {U} is the vector of nodal unknown variables (for instance displacement or water pressure), and {F} corresponds to the vector of applied nodal loads. Based on the two basic principles stated above, a finite element procedure consists of the following steps: (a) discretisation and selection of elements; (b) selection of the stress-strain relationships; (c) evaluation of element matrices; (d) assembly of elements matrices and introduction of boundary conditions; (e) solution to nodal unknowns; (f) computation of derived quantities, and analysis of results. (a) Discretisation and elements selection The elements used in conjunction with any discretisation process are selected with the view to obtaining sufficiently precise values of the nodal unknowns. Their shape and type depend therefore on the complexity of the boundaries of the discretised domain, and on the complexity of the FINITE ELEMENT MODELLING 715 physical behaviour to be modelled. In this respect, figure 14.1 illustrates the deformed shape of various one-, two-, and three-dimensional elements, each characterised by the number of nodes it contains. It is useful to mention at this stage that the number of variables at each node defines the number of degrees of freedom per node which are not necessarily identical for all nodes of every element in the mesh. Also, elements for which the same shape functions are used to define the unknown variable and the geometry of the element (i.e. its edges) are known as isoparametric elements. The (deformed) shapes are related to the type of the shape functions. Thus, linear elements are such that the unknowns vary linearly between any two connected nodes. The edges of a quadratic element, on the other hand, are curved since the corresponding shape functions consist of second-order polynomials. Cubic elements are generated using third-order polynomials. While the list of elements depicted in the figure is by no means exhaustive, it contains, however, some of the most widely used elements in finite element modelling in geotechnics. Thus, the three-noded triangle (element 4 in the figure) is the simplest two-dimensional element. Because of its associated linear shape functions, the strains (and consequently the stresses) are constant within the element. For that reason, the element is usually referred to as the constant strain triangle (CST). Although the CST is very easy to program, its use is limited to problems which do not involve the derivatives of the nodal variables such as computational fluid dynamics and seepage applications. Similarly, the same limitations apply to the four-noded quadrangle (element 7 in figure 14.1). For these reasons, the six-noded triangle and the eight-noded quadrangle (quadratic elements 5 and 8 respectively) are the two-dimensional isoparametric elements par excellence, although their performance depends on the integration rule used (see section 14.3). Also, the 15-noded quadratic (cubic strain) triangle (element 6 with an extra node on each side and three extra interior nodes, figure 14.1) is another quite widely used element which is now incorporated in many finite element packages. 716 FINITE ELEMENT MODELLING IN GEOTECHNICS LINEAR | QUADRATIC. | | CUBIC ELEMENTS | ELEMENTS —|_ ELEMENTS 2nodes Fnodes dnodes 1 / Z . rj 3 nodes | _ | 6 nodes 4 nodes 4 nodes 10 nodes Te nodes D \A\ 8 nodes Figure 14.1: Selective types of elements. FINITE ELEMENT MODELLING 717 These elements are particularly suitable for modelling plane strain and axisymmetric problems. Moreover, the quadratic nature of the associated shape functions makes it possible for domains with curved edges to be discretised in a precise way without the use of an excessive number of elements. It is perhaps worth mentioning that the shape functions associated with the eight-noded quadrangle are based on incomplete second-order polynomials (the straightforward mathematical details can be found in any book on finite elements) and, accordingly, the element is often referred to (in the jargon of finite elements) as a serendipity quadrangle. When it comes to three-dimensional modelling in geotechnics, the two most widely used elements are the ten-noded tetrahedron and the twenty-noded serendipity hexahedron (elements 11 and 14 in figure 14.1). These quadratic 3-D elements can be used successfully to estimate the displacement and stress fields. It must be borne in mind that the coding of mesh generation and output facilities for 3-D finite element analysis can be very complex. Accordingly, 3-D finite element modelling must only be used when neither plane strain nor axisymmetric conditions are suitable, or when the type of project justifies the cost of such a modelling (the case of the simulation of stresses and settlements beneath the foundation of a nuclear power plant for instance). It should be remembered that a finite element mesh consists generally of a combination of different types of compatible elements (i.e. nodes common to two elements must have the same degrees of freedom, and their shape functions must be characterised by polynomials of the same order). This might involve the use of transitional elements characterised by different shape functions on one side (to ensure a correct transition between elements) as illustrated in figure 14.2. More importantly, finite element modelling of soil structure interaction problems (examples include retaining, diaphragm or sheet-pile walls, laterally loaded piles, tunnels) may necessitate the use of interface elements to simulate the friction at the soil-structure interface (refer to figure 14.2). 718 FINITE ELEMENT MODELLING IN GEOTECHNICS common nodes quadratic element 1-D quadratic element (interface element) transitional element (variable nodes) Figure 14.2: Use of transitional and interface elements. (b) Selection of constitutive laws The predictions based on the outcome of a finite element modelling depend to a large extent on how realistic is the stress—strain relationship used in the calculations. It is a fact that most engineers prefer to use simple linear relationships, and in some cases related to homogeneous isotropic soils, such assumptions can be justified. However, for more complex soil conditions corresponding to layered soils with different stiffness characteristics, the assumption of a unique stress-strain relationship, let alone a linear one, can be markedly erroneous. It is essential to realise that finite element modelling, with a however refined mesh and sophisticated elements, will not per se offset any shortcomings related to the use of inappropriate stiffness parameters (that is stress-strain relationships), and as such, it cannot be emphasised enough that the numerical analyst must have a good feel, if not a good grasp, of soil behaviour prior to embarking on expensive numerical modelling. Non-linear stress-strain relationships usually lead to iterative procedures in computation, increasing noticeably the cost of calculations. Such cost must therefore be justified at least in terms of obtaining reliable predictions of stresses and deformations. (c) Evaluation of element matrices In finite element modelling, element matrices are first formulated in terms of local co-ordinates, so that for each element in the mesh, the element FINITE ELEMENT MODELLING 719 stiffness matrix and vector of nodal forces are established using an integral method such as the Galerkin weighted-residual method, for instance, thus yielding: t= iol : (14.2) try= [wendy [By being the transpose of the matrix [B] which is derived from the shape functions [NV], and [D] corresponds to the matrix of soil properties relating stresses and strains (refer to Hughes (1987), for example). The numerical integration of the element stiffness matrix is usually undertaken using the Gauss quadrature method, whose power is such that the integral in equation 14.2 needs only be evaluated at a few specific points known as Gauss integration points. It can be shown that using points of integration in each direction of the space, this method can integrate exactly up to a (2n—1)-order polynomial. For instance, it can readily be shown that in the case of the eight-noded serendipity quadrangle (clement 8 in figure 14.2), the integrand [B]’[D][B] in equation 14.2 is of the fourth order. Were the stiffness matrix for this element to be evaluated using nine integration points (that is 3 x3 so that m =3), an exact solution will be achieved since the use of n? =9 integration points can integrate exactly up to a fifth order polynomial (2n- 1 = 5). (d) Assembly of elements matrices and introduction of boundary conditions Once all elements' stiffness matrices are evaluated in local co-ordinates, a global stiffness matrix is then determined by first expanding each element matrix so that it is expressed in terms of all nodal variables of the entire mesh, then summing all matrices in a straightforward manner, yielding in the process the now familiar equation 14.1: [K]{U} = {F} 720 FINITE ELEMENT MODELLING IN GEOTECHNICS, in which the vector {U} contains all nodal variables (or degrees of freedom), and {F} represents the (expanded) vector containing all nodal forces. The boundary conditions in terms of nodal forces are inserted into the global vector {F'} in a straightforward way. However, a prescribed nodal displacement can be introduced in many ways. For example, were the displacement at node i to be prescribed at u; =5, then such a boundary condition can be satisfied if the diagonal coefficient in the stiffness matrix Kj; corresponding to node i is replaced by a large number (= 10° for example) and the quantity M8 is substituted for the nodal force F;. Notice that the global stiffness matrix is (usually) symmetric and has a dimension of (mx m) where m represents the total number of degrees of freedom contained in the entire mesh. For example a somewhat complex 2-D mesh for a tunnel having, say, 5000 nodes (which is not unusual) with only 2 degrees of freedom per node, yields a global stiffness matrix with a dimension of (10,000 x 10,000) = 108. Storing such a colossal matrix in its entirety would undoubtedly cause few problems; fortunately, only a few of the 100 million coefficients in this case are non-zero. What is more, the non-zero coefficients are situated in and around the diagonal of the matrix, thus forming a band. It is only the width of this band that is of interest, because its storage is done according to a dynamic allocation of space that allows the corresponding coefficients (which may include zeros) to be stored in lines so that they are easily retrieved when required. This method is known as the skyline storage method, the details of which can be found in Zienkiewicz and Taylor (1991), for instance. (e) Solution to nodal unknowns This step consists of calculating the vector of nodal unknowns {U} from the global equation: [K]{U} ={F} The obvious solution is to invert the stiffness matrix [K], so that: {U} =[K]'[F]. However, given the large size of [K], its inversion is not practical. A more effective solution consists of applying the triangular FINITE ELEMENT MODELLING 721 decomposition method in conjunction with the stiffness matrix, then using the forward elimination and back-substitution method to calculate the unknown nodal variables. Based on the Gaussian elimination method, it can be shown that, provided [K] is not singular (i.e. its determinant in not zero), it can always be written as the product of two matrices: [K]=[Z][S] (14.3) where [L] is a lower triangular matrix in which the diagonal coefficients are unity, and [S] is an upper triangular matrix, so that for example: 2 4 8 100]24 8 4 11 25 |=|2 10/039 6 18 46 3.2 1]0 04 [kK] = [] [s] Thus, substituting for [K] into the global relationship yields: [LIES 1{U} = fF} (14.4) which can then be solved in a very easy way using the forward elimination and back-substitution method (see Jennings and McKeown (1992), Hoffman (1992)) which consists of the following stages: - stage I: factorize [K]: [K]=[Z IIS] (14.5a) - stage 2: use a forward substitution to solve for {Z}: [L]{Z} = {F} (14.56) - stage 3: use a back-substitution to solve for {U}: [S]{U} = {Z} (14.5¢) Notice that {Z} in equations 14.5 represents a vector of intermediate variables. 722 FINITE ELEMENT MODELLING IN GEOTECHNICS Such a method of solution, however, may quickly become unsuitable for 3-D finite element analyses, both on grounds of storage capacity and the number of calculations. Under such circumstances, iterative solvers such as the one based on the pre-conditioned conjugate gradient method (see Jennings and McKeown (1992)) can be very useful in terms of efficiency and precision. (f) Calculation of derived quantities and analysis of results Once the primary unknowns in equation 14.4 are determined, the gradients or derived quantities can then be established. For instance, the strains are derived from displacements, and the stresses are calculated from the stress-strain constitutive relationships. Perhaps the most important step in a finite element modelling consists of analysing the outcome of computation. Once more, it is worth reiterating that the finite element method is only a method of solution and, as such, it cannot make for any deficiency related to soil stiffness parameters used in the calculations. It is naive to consider a result acceptable simply by virtue of the fact that it was generated through a finite element modelling. The engineer must be in a position to question the very raison d’étre of such a modelling in the first place. Is it needed ? If yes, then how can one get a reliable set of soil parameters (preferably measured in situ so as to minimise the effects of disturbance). Furthermore, it is essential that the soil-structure interaction mechanisms are well understood before embarking on time consuming expensive numerical modelling. It is also helpful if the working of the finite element method is understood, so that when it comes to the interpretation of results, any aberration can be attributed to its appropriate cause(s). In this respect, the following section contains some practical analyses related to the potential pitfalls of finite element modelling. 14.2 Effective stress analysis Thus far, the term stress has been used indiscriminately. However, there is often a need during a finite element analysis of geotechnical problems, to dissociate (numerically) effective stresses from porewater pressures EFFECTIVE STRESS ANALYSIS 723 according to the effective stress principle (an important aspect in geotechnics). Although the specific details of such an analysis is beyond the scope of the present text (reference can be made to Naylor (1974) for example), it is worth mentioning that an effective stress formulation can easily be incorporated in a finite element program. The procedure consists of using an expanded matrix of soil properties relating stresses and strains (matrix [D ] in equation 14.2) in the following manner (see Naylor et al. (1981): [D] = [D']+ [abn]? [Ky] (14.6a) {o} = {o/} + [m] {u} (14.65) where [m]” represents the transpose of the column matrix related to the effective stress principle, [D’] is the soil skeleton modulus matrix, and [Ky] corresponds to the bulk modulus of the pore fluid element. Note that all coefficients corresponding to shear stress components in [m] are zero, the remaining coefficients being equal to 1. All calculations involving element matrices, assembly and solution to the global system of equations thus formed are then undertaken as described previously. Once the strains are calculated, the ensuing stress increments are then evaluated using the following relationships: {Ao’} = [D/]f{Ae} (14.7a) {Au} = [Ky] {Aev} (14.76) {Au} being the excess (i.e. load induced) porewater pressure, and {Aey} represents the volumetric strain change. Notice that a drained analysis amounts to setting [Ky] to zero. More importantly, the effective stress analysis presented above does not apply to transient flow problems (i.e. consolidation problems). Also, it should be mentioned that the critical state model presented in chapter 6 can be (and is often) successfully coupled to finite element programs. In this respect, reference should be made to Britto and Gunn (1986) and Naylor et al. (1981). 724 FINITE ELEMENT MODELLING IN GEOTECHNICS 14.3 Finite element modelling of seepage and consolidation problems Although seepage problems are steady state problems, their finite element solution is nonetheless elaborated in a slightly different way from the one developed earlier, since pressure heads are the corresponding primary unknowns. Seepage problems are represented by an elliptic equation, the two- dimensional form of which has been established earlier (see equation 3.47, section 3.4): =0 (14.8) where ky and ky are the coefficients of soil permeability and 4 corresponds to the total head. Notice that for ky # ky, equation 14.8 is not a Laplace equation. Whilst developing a closed form solution to such an equation can be fraught with difficulties, mainly because in many cases of flow, the corresponding boundary conditions are very difficult to satisfy, the finite element method can be applied in a straightforward manner. It can be shown that, irrespective of the nature of flow (i.e. confined or unconfined), the discretisation of equation 14.8 invariably yields a global relationship: [K]{A} = {Q} (14.9) which is a familiar matrix relationship, since it is similar to equation 14.1. Here {/} is the vector of nodal variables, which in this case correspond to the total heads, and {Q} is the vector of nodal flow. It is interesting to mention that solving for total heads implies that each node has only one degree of freedom, in other words there is only one unknown quantity at each node. The global stiffness matrix [K] in equation 14.9 is obtained from the assembly of element matrices, each of which is calculated as follows: FINITE ELEMENT MODELLING: SEEPAGE AND CONSOLIDATION 725 [K]’= [ (By'tP IIB] av (14.10) where [B]’ corresponds to the transpose of matrix [B] which is derived from the shape functions, and [P] is the element permeability matrix (Naylor et al., 1981): ri-| § c| (14.11) Notice the diagonal nature of [P ]. The numerical solution to equation 14.9 depends on the boundary conditions and therefore on the nature of seepage. Thus, in the case of confined flow as illustrated in figure 14.3 (refer also to section 3.4 for more details related to this type of flow), the corresponding boundary conditions are: - along AB: H=h, +d, -alongCD: H=hy+d aH. - along EF: ape 0 (impermeable side) Figure 14.3: Boundary conditions related to confined flow. 726 FINITE ELEMENT MODELLING IN GEOTECHNICS For unconfined flow, on the other hand (see figure 14.4), the boundary conditions are: - along AB: H=h, -along BC: H=y and u=0 (u being the porewater pressure, tefer to section 3.4) - along AC: = =0 (impermeable side) in both cases, H is the total head expressed with respect to the datum represented by the impermeable base. The free surface in the case of unconfined flow presents an additional problem as its location is not known a priori. The precise location can be found using iterative techniques such as the one suggested by Smith and Griffiths (1998), that consists of deforming the mesh until, eventually, its upper surface coincides with the free surface characterised by the boundary condition of a total head identical to the elevation head at any given point. B Figure 14.4: Boundary conditions related to unconfined flow. Time-dependent problems such as consolidation are represented by a parabolic equation (refer to the consolidation equation 4.14, section 4.3): @u_ du eve =a (14.12) for which it can readily be shown that finite clement modelling yields the following relationship: te] 2] + Kw} =F} (14.13) PRACTICAL ASPECTS OF FINITE ELEMENT MODELLING 727 where [C ] represents the matrix of soil consolidation characteristics. The nodal variables {U} this time correspond to the porewater pressure; the remaining symbols having the same meaning as in equation 14.1. A solution to equation 14.13 can be obtained through a combination of finite difference formulation (to discretise time) and finite elements. Consequently, using a forward difference operator: (22) = LE hgh] 14.14) (At being the time increment), then inserting the latter quantity into equation 14.13 and rearranging, the following algorithm can easily be established: Way = ater i},+(Licl-ta)tv},] a4) which can then be solved on a step-by-step basis. 14.4 Practical aspects of finite element modelling in geotechnics Finite element modelling of soil-structure interaction problems needs to be planned and undertaken carefully, so that any anomalies in the results can be spotted and remedied. The first step consists of discretising the domain, in other words generating a mesh. Although some finite element design programs contain mesh-generation pre-processors, it must be remembered that a mesh should always constitute a compromise between the computer capacity to store data and execute calculations (this capacity can, in some cases, be exceeded even by the standard of very powerful machines), and reasonable predictions in terms of outcome such as strains, stresses, porewater pressure etc. Thus, a reasonable mesh should be refined (i.e. formed of smaller elements) near any applied load where the stress and displacement gradients are expected to be large, and around any geometric or material discontinuities (or singularities to use a finite element jargon) such as 728 FINITE ELEMENT MODELLING IN GEOTECHNICS tunnel openings or excavations. Also, the mesh needs to be refined around the areas where changes in material properties occur, as in the case of different soil layers, or around a soil-structure interface (pile-soil, retaining structure-soil, tunnel-soil etc.). In such cases, interface elements may be needed to simulate friction at the soil-structure interface, particularly when the structure surface is relatively smooth. Generally, the size of elements used in a mesh depends on the loading conditions and geometric discontinuities stated above. However, numerical evidence seems to indicate that the aspect ratio of an element (i.e. the ratio of the Jargest to the smallest dimensions of an element) must be kept within reasonable limits. In this respect, an aspect ratio smaller than 3 ensures satisfactory results in terms of stresses, unless the soil behaviour is markedly non-linear, in which case an even smaller ratio is required. As far as the shape of elements is concerned, and whenever practicable, triangles, quadrangles and hexahedra used to generate a mesh should be as near as possible to equilateral triangles, squares and cubes respectively, so as to avoid any excessive distortion of these elements in the advent of large displacements. Notice that some sophisticated finite element programs offer the possibility of automatically regenerating the mesh while taking into account the level of deformation of different elements. The size of elements should be increased progressively around the refined portions of the mesh in a way that ensures a smooth transition from small to larger size elements. This can be achieved if the ratio of the areas (or volumes in 3-D) of two adjacent elements does not exceed 2. In any case, it is strongly advised (perhaps one should say it is logical) to avoid using distorted elements such as the ones illustrated in figure 14.5 when generating a mesh. The reason being that, once loaded, each of these elements can potentially deform to the point where the stiffness matrix becomes singular (i.e. with a zero determinant, so that it cannot be inverted) thus causing numerical instability. As a guideline, any angle a within an element must be such that 15° < a < 165°. Also, the middle node in a quadratic element should be situated within the middle third of the side [see figure 14.5] (Zienkiewicz and Taylor, 1991). Transitional elements (refer to figure 14.2 for the principle involved) are mainly used in conjunction with 3-D modelling. PRACTICAL ASPECTS OF FINITE ELEMENT MODELLING 729 Figure 14.5: Unacceptable elements. The type of elements used in a finite element modelling depends on the type of problem to be analysed. Nevertheless, one has to bear in mind that the strains and stresses are derived from displacements and, as such, using linear elements (ie. elements characterised by linear shape functions) automatically yields constant strains and stresses within each element, which can be markedly erroneous, particularly in relation to stress distribution. Accordingly, the use of linear elements (refer to figure 14.1) is not advised. Quadratic elements on the other hand, especially the six-noded triangle and the eight-noded quadrangle, are widely used to model different type of problems in geomechanics. The calculation of strains and stresses is undertaken within each element at specific integration points. Often, the integration points used to calculate the strains (and hence the stresses) are fewer than those used to calculate the element stiffness matrix. For example, whilst the stiffness matrix of the eight-noded serendipity quadrangle is calculated using 3 x3 integration points, numerical evidence shows that, in most cases, satisfactory results in terms of displacements are obtained from a reduced integration using 2 x 2 integration points. In fact, in this case, an exact integration using 3 x3 points yields an over-stiff response, so underestimating displacements. Situations might arise whereby a mechanism can occur as a result of a 730 FINITE ELEMENT MODELLING IN GEOTECHNICS reduced integration using 2 x2 points; however, these cases are extreme. As regards the six-noded triangle, a happy balance can be achieved using three (non-Gauss) integration points. On the other hand the use of four integration points in conjunction with the ten-noded tetrahedron (one of the most widely used 3-D elements) provides satisfactory results in terms of displacements. Interested readers can refer to the paper by Naylor (1994) on integration rules. Once calculated at the integration points, the stresses are then extrapolated to the element nodes and, because each element has its own stiffness matrix, the magnitude of stresses resulting from each element at a common node such as the one depicted in figure 14.6 is generally different. This stress discontinuity can be smoothed by taking the average value at the node, so that the corresponding nodal stress results from the contribution of all adjacent elements. However, nodal averaging should not be used for stress and strain calculations at nodes connecting elements with different stiffnesses (for example elements corresponding to different soil layers). Such boundaries are characterised by a stress discontinuity, except for the stress component normal to the boundary. Numerical evidence shows that the stress field is not affected by the stress discontinuity provided that a refined mesh with smaller size elements around theses boundaries is used. Figure 14.6; Stress calculation at a common node. When using finite element modelling, special attention must be paid to the degree and type of anisotropy of the soil-structure material. For example the distribution with depth of the vertical stress generated by a uniform surface load within a thick homogeneous isotropic soil layer is markedly different in shape and in magnitude from that generated by the same load within a layered anisotropic soil. PRACTICAL ASPECTS OF A FINITE ELEMENT MESH 731 As mentioned earlier, geometric or material discontinuities constitute singularities which are unavoidable for a large number of soil-structure interaction problems. This can generate inaccurate stress fields near the singularities (for example, an infinite stress beneath a point load is computed as a finite quantity). However, these effects can be minimised by using a refined mesh around the singularity, so that a right-angle boundary, for instance, can be transformed into a smooth curved boundary. 14.5 Practical aspects of a finite element mesh related to foundations A finite element mesh related to a given problem in geomechanics must always take into account different aspects evoked previously and linked to elements type and size, and particularly to the nature of the problem (axisymmetry, anisotropy, drainage conditions, soil-structure interaction, geometric singularities, and nature of loading). Consequently, the following guidelines should be viewed with a sufficient degree of flexibility, so that local conditions are appropriately considered. As such, they should in no way be regarded as rules. Based on the previous analysis, a finite element (axisymmetric) mesh in the case of a shallow foundation on an isotropic homogeneous soil, with a width B usually includes an area extending to about 5B laterally and 8B vertically as illustrated in figure 14.7, an area within which most of the stress variations are expected to occur (refer to figures 2.37-2.40, section 2.5). The conditions imposed at the mesh edges in figure 14.7 allow for a vertical movement (i.e. u=0) along the vertical boundaries, while restricting any movement to u=v=0 at the bottom horizontal boundary, where the stresses (and therefore the displacements) are expected to decay. Also, notice how the mesh is refined beneath and around the foundation, with increasing elements size as one moves away from the foundation in each direction. Were the soil conditions to be different as in the case, for example, of ansiotropic multi-layered and partially submerged soils, then the overall size of mesh as well as the size and type of elements would need to be altered drastically so as to reflect the markedly different characteristics of 732 FINITE ELEMENT MODELLING IN GEOTECHNICS the problem. Hence the informative nature of the dimensions indicated in figure 14.7. 2 5B [ae shallow foundation u=0 ~ 8B u=v=0 Figure 14.7: Typical (axisymmetric) mesh dimensions for anisolated shallow foundation on isotropic homogeneous soil. For a deep pile foundation, the mesh depends on the type of loading and the type of pile. Thus, for a single axially loaded pile with a length L embedded in an isotropic homogeneous soil, the mesh should ideally cover an area 3L-deep and about 30-pile diameters wide as depicted in figure 14.8. The typical (axisymmetric) mesh in the figure consists mainly of eight-noded quadrangles and a few six-noded triangles. The size of the elements reflects the stress distribution generated around the pile shaft and tip depicted in figure 9.31 (refer to section 9.5). The boundary conditions are of a similar type to the ones described previously in the case of shallow foundations. However, in this case, there is a need to use interface elements in order to simulate the friction developed along the pile shaft. It is interesting to notice that the mesh in figure 14.8 is used to simulate the behaviour of an already embedded pile. The numerical simulation of pile PRACTICAL ASPECTS OF A FINITE ELEMENT MESH 733 driving is very complex in nature, not least because soil failure around the pile shaft and tip has to occur every time the pile is struck with the driving tool, so that it can be driven to the required depth, or sometimes until the occurrence of a refusal. Dye ~ 30d w2L Figure 14.8: Typical (axisymmetric) mesh dimensions for a single axially loaded pile (isotropic homogeneous soil conditions). A laterally loaded single pile, on the other hand, is characterised by a markedly different behaviour. This type of loading occurs especially in conjunction with bridges, flyovers and retaining structures founded on piles, and can be applied either actively or passively as depicted in figure 14.9. In both cases, the solution in terms of displacement, bending moment, shear stress and lateral pressure distribution with depth depends on the soil-pile interaction, in other words on both soil and pile stiffness characteristics. The soil—pile interaction is depicted in figure 14.10 whereby the resistance to the active load (a combination of a horizontal load and a bending moment applied at the pile head) is provided by the pile stiffness and the 734 FINITE ELEMENT MODELLING IN GEOTECHNICS soil reaction per unit length (that is, the force per unit length induced within the soil mass by the active load). active loading passive loading He (Ty Me Figure 14.9: Active and passive lateral loading of piles. While the pile stiffness can be assumed to be constant along the pile shaft, the soil stiffness, on the other hand, depends generally on the depth of embedment. Figure 14.10 shows that at a given depth z, the relationship between the soil reaction P and the lateral displacement y (i.e. the pile deflection) is non-linear. Furthermore, the nature of the relationship varies with depth. Figure 14.10: Soil-structure interaction along the shaft of an actively loaded single pile. PRACTICAL ASPECTS OF A FINITE ELEMENT MESH 735 The profiles of pile deflection, bending moment and soil reaction are also affected by the pile dimensions and the boundary conditions: a long slender pile behaves in a different way from a short rigid pile as illustrated in figures 14.11 and 14.12 in the case of actively loaded piles embedded in an isotropic homogeneous sand. The figures also depict the difference in behaviour between a free head and a fixed head pile. Hy, 2 Mo soil reaction bending moment Figure 14.11: Profiles of deflection, soil reaction and bending moment Sor a long pile in a cohesionless soil. Hy . AES Mo soil reaction bending ‘moment Figure 14.12: Profiles of deflection, soil reaction and bending moment for a short pile in a cohesionless soil. Accordingly, a finite element mesh in the case of a laterally loaded pile should reflect the soil, pile and loading conditions. Because of the nature of the problem, no axisymmetry can be applied, and the mesh dimensions indicated in figure 14.13 should be applied sensibly. 736 FINITE ELEMENT MODELLING IN GEOTECHNICS Figure 14.13: Typical mesh dimensions for a single laterally loaded pile (isotropic homogeneous soil conditions). A pilegroup is much more difficult to model, especially in the absence of axisymmetry (which is generally the case). The difficulties are further illustrated by the fact that several modes of failure can potentially develop (a single pile failure within the group, a row of piles failing at the same time or a block failure en masse, refer to chapter 9). Under such circumstances, a three-dimensional mesh might become a necessity, with all the consequences elicited earlier concerning the potentially very large size of global matrices. However, notwithstanding the somewhat high cost that might be incurred in some cases, 3-D modelling of pilegroups is now becoming an almost routine operation. Whenever the 3-D problem can be reduced to an axisymmetric problem, then a mesh such as the typical one depicted in figure 14.14 corresponding to a pilegroup embedded in an isotropic soil can be used. The dimensions of the area covered by the mesh extend to about 3Z in depth, and at least a distance Z outside the pile situated at the group edge. Such a mesh should reflect the stress distribution generated within the soil mass by a pilegroup. PRACTICAL ASPECTS OF A FINITE ELEMENT MESH 737 u=v=0 Figure 14.14: Typical (axisymmetric) mesh dimensions ‘for an axially loaded pilegroup in an isotropic soil. 14.6 Finite element mesh related to embankments, retaining structures and tunnels The finite element modelling of soil loading through an embankment with a height H must be such that the mesh area covers most of the stress increase generated within the soil mass. Also, the area must take into account the possibility of a long term or a short term slope stability failure. Accordingly, for an isotropic homogeneous soil layer, the 2-D mesh should be characterised by a lateral dimension of at least four times the embankment length, and a minimum depth of 5H or the depth of stiff substratum, whichever is smaller (see figure 14.15). The mesh should be refined in regions where the maximum stress generated by the loading is expected, and the size of elements should be increased gradually as one moves away from these regions. Note that were the depth of the stiff substratum to apply, then the boundary condition «= v=0 indicated in the figure would correspond to a relatively small change in stiffness between the two layers. If the bottom layer were a rock, for example, overlain by a soft clay, then the boundary condition should be changed to allow for a lateral displacement, in which case v = 0. 738 FINITE ELEMENT MODELLING IN GEOTECHNICS >3L — embankment u=0 i Figure 14.15: Typical mesh dimensions for an isotropic homogeneous soil loaded through an embankment. As regards a retaining wall with height H and base width L (see figure 14.16), the mesh should be wide enough to include not only the stress changes in the soil mass beneath the wall, but also the potential development of long term active and passive stress failures, as well as the possibility of a deep circular failure. Consequently, for a wall retaining an isotropic homogeneous soil, it is advised to use a mesh with a minimum lateral dimension of 2H in front of the wall and 3H behind the wall, depending on the type of soil. The mesh should extend a minimum depth of 6Z or the depth of the stiff substratum, whichever is smaller as illustrated in figure 14.16. Also, interface elements may have to be used if the wall surface is relatively smooth. The boundary condition related to the stiff substratum indicated in the figure reflects the substantial change in stiffness between the two layers of soil. PRACTICAL ASPECTS OF A FINITE ELEMENT MESH 739 retaining wall 7 i | u=0 Figure 14.15: Typical mesh dimensions for a wall retaining an isotropic homogeneous soil. Similar guidelines apply to diaphragm and sheet-pile walls (refer to figure 14.17). Finite element modelling of an underground cavity or a tunnel requires an engineering judgement as to how such a task can best be undertaken cost-effectively. Thus 3-D modelling is appropriate when there is a variation of soil stratification along the tunnel axis, or when stress conditions near the tunnel face need be studied. On the other hand, an axisymmetrical analysis is suitable in the case of a vertical shaft when the soil stratification varies in the vertical direction only. There are instances in which plane deformations can be assumed to prevail, and therefore only a cross-section of the tunnel needs to be discretised. Under such circumstances, the area covered by the mesh must reflect the expected stress distribution and deformations around the opening. 740 FINITE ELEMENT MODELLING IN GEOTECHNICS diaphragm wall ~3H (or sheet pile) ee Figure 14.17: Typical mesh dimensions for a diaphragm wall retaining an isotropic homogeneous soil. As explained earlier in section 12.2, the stress arching that occurs around the opening is related to its depth and, accordingly, the dimensions of the mesh area should typically extend to a depth of around Sd beneath the tunnel invert, and should include the entire height of soil above the crown, unless the cover exceeds 10d, in which case the soil thickness above the crown can be limited to 5d (Mestat, 1997). Laterally, the dimension of the mesh area should be extended to about 6d from the tunnel axis as illustrated in figures 14.18(a) and (6). Obviously, these dimensions only represent a guideline, and the engineer must always seek a compromise between reliable predictions in terms of stress-strain distribution, and cost. In so doing, the dimensions advocated above might need to be adjusted to suite the site conditions. PRACTICAL ASPECTS OF A FINITE ELEMENT MESH TAL up to 10d [eee i ae v=0 . 3 | tigen Figure 14.18: Typical mesh dimensions for (a) a shallow opening and (b) a deep tunnel (from Mestat (1997), by permission). 742 FINITE ELEMENT MODELLING IN GEOTECHNICS References Bathe, K. J. (1982) Finite Element Procedures in Engineering Analysis. Prentice-Hall, Englewood Cliffs, New Jersey. Bathe, K. J. (1996) Numerical Methods in Finite Element Analysis, 3rd edn. Prentice-Hall, Englewood Cliffs, New Jersey. Britto, A. and Gunn, M. J. (1986) Critical State Soil Mechanics via Finite Elements. Ellis Horwood, Chichester. Griffiths, D. V. and Smith, I. M. (1991) Numerical Methods for Engineers. Blackwell, Oxford. Hoffman, J. D. (1992) Numerical Methods for Engineers and Scientists. McGraw-Hill, New York. Hughes, T. J. R. (1987) The Finite Element Method: Linear, Static and Dynamic Finite Element Analysis. Prentice-Hall, Englewood Cliffs, New Jersey. Jennings, A. and McKeown, J. J. (1992) Matrix computation, 2nd edn. John Wiley & Sons, Chichester. Mestat, P. (1997) Maillage d’élements finis pour les ouvrages de géotechnique: conseils et recommandations. Bulletin de Laboratoires des Ponts et Chaussées, 212, pp. 39-64. Naylor, D. J. (1974) Stresses in nearly incompressible materials for finite elements with application to the calculation of excess pore pressures. International Journal of Numerical Methods in Engineering, 8, pp. 443-460. Naylor, D. J. (1994) On integration rules for triangles. Proceedings of the 3rd European Conference on Numerical Methods in Geotechnical Engineering, Manchester (ed. I. M. Smith), Balkema, Rotterdam. Naylor, D. J., Pande, G. N., Simpson, B. and Tabb, R. (1981) Finite Elements in Geotechnical Engineering. Pineridge Press, Swansea, UK. Owen, D. R. J. and Hinton, E. (1980) Finite Elements in Plasticity: Theory and Practice. Pineridge Press, Swansea. Reddy, J. N. (1993) An Introduction to the Finite Element Method. McGraw-Hill, New York. Smith, I. M. and Griffiths, D. V. (1998) Programming the Finite Element Method, 3rd edn. John Wiley & Sons, New York. Stasa, F. L. (1985) Applied Finite Element Analysis for Engineers. CBS International Edition, New York. Zienkiewicz, O. C. and Taylor, R. (1991) The Finite Element Method, (2 volumes) 4th edn. McGraw-Hill, London.

Vous aimerez peut-être aussi